State-transition matrix

In control theory, the state-transition matrix is a matrix whose product with the state vector x {\displaystyle x} at an initial time t 0 {\displaystyle t_{0}} gives x {\displaystyle x} at a later time t {\displaystyle t} . The state-transition matrix can be used to obtain the general solution of linear dynamical systems.

Linear systems solutions

The state-transition matrix is used to find the solution to a general state-space representation of a linear system in the following form

x ˙ ( t ) = A ( t ) x ( t ) + B ( t ) u ( t ) , x ( t 0 ) = x 0 {\displaystyle {\dot {\mathbf {x} }}(t)=\mathbf {A} (t)\mathbf {x} (t)+\mathbf {B} (t)\mathbf {u} (t),\;\mathbf {x} (t_{0})=\mathbf {x} _{0}} ,

where x ( t ) {\displaystyle \mathbf {x} (t)} are the states of the system, u ( t ) {\displaystyle \mathbf {u} (t)} is the input signal, A ( t ) {\displaystyle \mathbf {A} (t)} and B ( t ) {\displaystyle \mathbf {B} (t)} are matrix functions, and x 0 {\displaystyle \mathbf {x} _{0}} is the initial condition at t 0 {\displaystyle t_{0}} . Using the state-transition matrix Φ ( t , τ ) {\displaystyle \mathbf {\Phi } (t,\tau )} , the solution is given by:[1][2]

x ( t ) = Φ ( t , t 0 ) x ( t 0 ) + t 0 t Φ ( t , τ ) B ( τ ) u ( τ ) d τ {\displaystyle \mathbf {x} (t)=\mathbf {\Phi } (t,t_{0})\mathbf {x} (t_{0})+\int _{t_{0}}^{t}\mathbf {\Phi } (t,\tau )\mathbf {B} (\tau )\mathbf {u} (\tau )d\tau }

The first term is known as the zero-input response and represents how the system's state would evolve in the absence of any input. The second term is known as the zero-state response and defines how the inputs impact the system.

Peano–Baker series

The most general transition matrix is given by the Peano–Baker series

Φ ( t , τ ) = I + τ t A ( σ 1 ) d σ 1 + τ t A ( σ 1 ) τ σ 1 A ( σ 2 ) d σ 2 d σ 1 + τ t A ( σ 1 ) τ σ 1 A ( σ 2 ) τ σ 2 A ( σ 3 ) d σ 3 d σ 2 d σ 1 + . . . {\displaystyle \mathbf {\Phi } (t,\tau )=\mathbf {I} +\int _{\tau }^{t}\mathbf {A} (\sigma _{1})\,d\sigma _{1}+\int _{\tau }^{t}\mathbf {A} (\sigma _{1})\int _{\tau }^{\sigma _{1}}\mathbf {A} (\sigma _{2})\,d\sigma _{2}\,d\sigma _{1}+\int _{\tau }^{t}\mathbf {A} (\sigma _{1})\int _{\tau }^{\sigma _{1}}\mathbf {A} (\sigma _{2})\int _{\tau }^{\sigma _{2}}\mathbf {A} (\sigma _{3})\,d\sigma _{3}\,d\sigma _{2}\,d\sigma _{1}+...}

where I {\displaystyle \mathbf {I} } is the identity matrix. This matrix converges uniformly and absolutely to a solution that exists and is unique.[2]

Other properties

The state transition matrix Φ {\displaystyle \mathbf {\Phi } } satisfies the following relationships:

1. It is continuous and has continuous derivatives.

2, It is never singular; in fact Φ 1 ( t , τ ) = Φ ( τ , t ) {\displaystyle \mathbf {\Phi } ^{-1}(t,\tau )=\mathbf {\Phi } (\tau ,t)} and Φ 1 ( t , τ ) Φ ( t , τ ) = I {\displaystyle \mathbf {\Phi } ^{-1}(t,\tau )\mathbf {\Phi } (t,\tau )=I} , where I {\displaystyle I} is the identity matrix.

3. Φ ( t , t ) = I {\displaystyle \mathbf {\Phi } (t,t)=I} for all t {\displaystyle t} .[3]

4. Φ ( t 2 , t 1 ) Φ ( t 1 , t 0 ) = Φ ( t 2 , t 0 ) {\displaystyle \mathbf {\Phi } (t_{2},t_{1})\mathbf {\Phi } (t_{1},t_{0})=\mathbf {\Phi } (t_{2},t_{0})} for all t 0 t 1 t 2 {\displaystyle t_{0}\leq t_{1}\leq t_{2}} .

5. It satisfies the differential equation Φ ( t , t 0 ) t = A ( t ) Φ ( t , t 0 ) {\displaystyle {\frac {\partial \mathbf {\Phi } (t,t_{0})}{\partial t}}=\mathbf {A} (t)\mathbf {\Phi } (t,t_{0})} with initial conditions Φ ( t 0 , t 0 ) = I {\displaystyle \mathbf {\Phi } (t_{0},t_{0})=I} .

6. The state-transition matrix Φ ( t , τ ) {\displaystyle \mathbf {\Phi } (t,\tau )} , given by

Φ ( t , τ ) U ( t ) U 1 ( τ ) {\displaystyle \mathbf {\Phi } (t,\tau )\equiv \mathbf {U} (t)\mathbf {U} ^{-1}(\tau )}

where the n × n {\displaystyle n\times n} matrix U ( t ) {\displaystyle \mathbf {U} (t)} is the fundamental solution matrix that satisfies

U ˙ ( t ) = A ( t ) U ( t ) {\displaystyle {\dot {\mathbf {U} }}(t)=\mathbf {A} (t)\mathbf {U} (t)} with initial condition U ( t 0 ) = I {\displaystyle \mathbf {U} (t_{0})=I} .

7. Given the state x ( τ ) {\displaystyle \mathbf {x} (\tau )} at any time τ {\displaystyle \tau } , the state at any other time t {\displaystyle t} is given by the mapping

x ( t ) = Φ ( t , τ ) x ( τ ) {\displaystyle \mathbf {x} (t)=\mathbf {\Phi } (t,\tau )\mathbf {x} (\tau )}

Estimation of the state-transition matrix

In the time-invariant case, we can define Φ {\displaystyle \mathbf {\Phi } } , using the matrix exponential, as Φ ( t , t 0 ) = e A ( t t 0 ) {\displaystyle \mathbf {\Phi } (t,t_{0})=e^{\mathbf {A} (t-t_{0})}} . [4]

In the time-variant case, the state-transition matrix Φ ( t , t 0 ) {\displaystyle \mathbf {\Phi } (t,t_{0})} can be estimated from the solutions of the differential equation u ˙ ( t ) = A ( t ) u ( t ) {\displaystyle {\dot {\mathbf {u} }}(t)=\mathbf {A} (t)\mathbf {u} (t)} with initial conditions u ( t 0 ) {\displaystyle \mathbf {u} (t_{0})} given by [ 1 ,   0 ,   ,   0 ] T {\displaystyle [1,\ 0,\ \ldots ,\ 0]^{T}} , [ 0 ,   1 ,   ,   0 ] T {\displaystyle [0,\ 1,\ \ldots ,\ 0]^{T}} , ..., [ 0 ,   0 ,   ,   1 ] T {\displaystyle [0,\ 0,\ \ldots ,\ 1]^{T}} . The corresponding solutions provide the n {\displaystyle n} columns of matrix Φ ( t , t 0 ) {\displaystyle \mathbf {\Phi } (t,t_{0})} . Now, from property 4, Φ ( t , τ ) = Φ ( t , t 0 ) Φ ( τ , t 0 ) 1 {\displaystyle \mathbf {\Phi } (t,\tau )=\mathbf {\Phi } (t,t_{0})\mathbf {\Phi } (\tau ,t_{0})^{-1}} for all t 0 τ t {\displaystyle t_{0}\leq \tau \leq t} . The state-transition matrix must be determined before analysis on the time-varying solution can continue.

See also

  • Magnus expansion
  • Liouville's formula

References

  1. ^ Baake, Michael; Schlaegel, Ulrike (2011). "The Peano Baker Series". Proceedings of the Steklov Institute of Mathematics. 275: 155–159. doi:10.1134/S0081543811080098. S2CID 119133539.
  2. ^ a b Rugh, Wilson (1996). Linear System Theory. Upper Saddle River, NJ: Prentice Hall. ISBN 0-13-441205-2.
  3. ^ Brockett, Roger W. (1970). Finite Dimensional Linear Systems. John Wiley & Sons. ISBN 978-0-471-10585-5.
  4. ^ Reyneke, Pieter V. (2012). "Polynomial Filtering: To any degree on irregularly sampled data". Automatika. 53 (4): 382–397. doi:10.7305/automatika.53-4.248. hdl:2263/21017. S2CID 40282943.

Further reading

  • Baake, M.; Schlaegel, U. (2011). "The Peano Baker Series". Proceedings of the Steklov Institute of Mathematics. 275: 155–159. doi:10.1134/S0081543811080098. S2CID 119133539.
  • Brogan, W.L. (1991). Modern Control Theory. Prentice Hall. ISBN 0-13-589763-7.
Wikibooks has a book on the topic of: Control Systems/Time Variant System Solutions
  • v
  • t
  • e
Matrix classes
Explicitly constrained entriesConstantConditions on eigenvalues or eigenvectorsSatisfying conditions on products or inversesWith specific applicationsUsed in statisticsUsed in graph theoryUsed in science and engineering
Related terms